Aminoboranediyl-Bridged Zirconocenes: Highly Active Olefin

Clinton L. Lund , Samuel S. Hanson , Gabriele Schatte , J. Wilson Quail , and Jens Müller. Organometallics 2010 29 (22), 6038-6044. Abstract | Full T...
13 downloads 0 Views 84KB Size
2288

Organometallics 1999, 18, 2288-2290

Aminoboranediyl-Bridged Zirconocenes: Highly Active Olefin Polymerization Catalysts Arthur J. Ashe, III,* Xinggao Fang, and Jeff W. Kampf Department of Chemistry, The University of Michigan, Ann Arbor, Michigan 48109-1055 Received March 12, 1999 Summary: Novel aminoboranediyl-bridged zirconocenes [i-Pr2NB(η5-C5H4)2]ZrCl2 (8b) and [i-Pr2NB(η5-1-C9H8)2]ZrCl2 (4b) have been prepared and structurally characterized. When activated by excess methylaluminoxane, 4b and 8b form highly active catalysts for the polymerization of ethylene. The catalyst derived from 4b converts propylene to isotactic polypropylene. The Brintzinger type of chiral group 4 bridged metallocenes (11 and 22) have been widely used as stereoselective homogeneous catalysts for olefin polymerization3 and in organic synthesis.4 Although a number of different bridging groups have been studied,5 the CH2CH2 and SiMe2 bridges are most commonly used.6 Bridges by first-row elements other than carbon are particularly rare. Boron-bridged 3b7 and a related bis Cp complex 5b8 have been reported by Rufanov, Lorberth, and co-workers, while Shapiro et al. have examined the coordinated boron-bridged 6b.9 Compounds 3b (1) (a) Wild, F. R. W. P.; Zsolnai, L.; Huttner, G.; Brintzinger, H. H. J. Organomet. Chem. 1982, 232, 233. (b) Wild, F. R. W. P.; Wasiucionek, M.; Huttner, G.; Brintzinger, H. H. J. Organomet. Chem. 1985, 288, 63. (c) Kaminsky, W.; Ku¨lper, K.; Brintzinger, H. H.; Wild, F. W. R. P. Angew. Chem., Int. Ed. Engl. 1985, 24, 507. (2) (a) Herrmann, W. A.; Rohrmann, J.; Herdtweck, E.; Spaleck, W.; Winter, A. Angew. Chem., Int. Ed. Engl. 1989, 28, 1511. (b) Spaleck, W.; Antberg, M.; Rohrmann, J.; Winter, A.; Bachmann, B.; Kiprof, P.; Behm, J.; Herrmann, W. A. Angew. Chem., Int. Ed. Engl. 1992, 31, 1347. (3) For recent reviews, see: (a) Brintzinger, H. H.; Fischer, D.; Mu¨lhaupt, R.; Rieger, B.; Waymouth, R. M. Angew. Chem., Int. Ed. Engl. 1995, 34, 1143. (b) Ziegler catalysts: Fink, G., Mu¨lhaupt, R., Brintzinger, H. H., Eds.; Springer-Verlag: Berlin, Germany, 1995. (c) Coates, G. W.; Waymouth, R. M. In Comprehensive Organometallic Chemistry II; Vol. 12 (Volume Ed.: Hegedus, L. S.); Pergamon: Oxford, U.K., 1995; p 1193. (d) Aulbach, M.; Ku¨ber, F. Chem. Unserer Zeit 1994, 28, 197. (4) Hoveyda, A. H.; Morken, J. P. Angew. Chem., Int. Ed. Engl. 1996, 35, 1263. (5) (a) Ge: Ko¨pf, H.; Kahl, W. J. Organomet. Chem. 1974, 64, C37. Klouras, N. Monatsch. Chem. 1981, 112, 887. Ko¨pf, H.; Klouras, N. Z. Naturforsch. 1983, 38B, 34. Chen, Y.-X.; Rausch, M. D.; Chien, J. C. W. Organometallics 1994, 13, 748. Schumann, H.; Esser, L.; Loebel, J.; Dietrich, A.; van der Helm, D.; Ji, X. Organometallics 1991, 10, 2585. (b) Sn: Herrmann, W. A.; Morawietz, M. J. A.; Herrmann, H.F.; Ku¨ber, F. J. Organomet. Chem. 1996, 509 115. (c) P: Ko¨pf, H.; Klouras, N. Monatsh. Chem. 1983, 114, 243. (d) As: Klouras, N. Z. Naturforsh. 1991, 46B, 647. Anderson, G. K.; Lin, M. Organometallics 1988, 7, 2285; Inorg. Chim. Acta 1988, 142, 7. Schaverien, C. J.; Ernst, R.; Terlouw, W.; Schut, P.; Sudmeijer, O.; Budzelaar, P. H. M. J. Mol. Catal., A-Chem. 1998, 128, 245. Leyser, N.; Schmidt, K.; Brintzinger, H.-H. Organometallics 1998, 17, 2155. Shin, J. H.; Hascall, T.; Parkin, G. Organometallics 1999, 18, 6. (6) (a) Halterman, R. L. Chem. Rev. 1992, 92, 965. (b) Ryan, E. J. In Comprehensive Organometallic Chemistry II, Vol. 4 (Volume Ed.: Lappert, M. F.); Pergamon: Oxford, U.K., 1995; p 483. (7) Rufanov, K.; Avtomonov, E.; Kazennova, N.; Kotov, V.; Khvorost, A.; Lemenovskii, D.; Lorberth, J. J. Organomet. Chem. 1997, 536537, 361. (8) Rufanov, K. A.; Kotov, V. V.; Kazennova, N. B.; Lemenovskii, D. A.; Avtomonov, E. V.; Lorberth, J. J. Organomet. Chem. 1996, 525, 287. (9) Stelck, D. S.; Shapiro, P. J.; Basickes, N.; Rheingold, A. L. Organometallics 1997, 16, 4546.

and 5b are rather labile and were poorly characterized, while the relatively more stable 6b (L ) PMe3) showed only a low activity toward polymerization of ethylene. Very recently Braunschweig et al. have prepared and structurally characterized the BN(SiMe3)2-bridged titanium complex 7a.10,11

Our prior work on olefin polymerization using boratabenzene zirconium complexes 912 and 1013 showed that a change of the B-substituent from Ph to N(i-Pr)2 markedly effected their polymerization activities.14 Thus we felt that the use of aminoborane bridges might result in ansa-zirconocenes with novel properties. Specifically, the aminoboron bridge should (1) mask the electrophilicity of boron by B-N π-bonding, (2) offer a more rigid bridge, and (3) result in an electron-rich boron which should diminish the electrophilicity of Zr. On the basis of these considerations we have prepared and structurally characterized 4b and 8b. On activation by excess methylaluminoxane (MAO), these complexes give olefin polymerization catalysts with remarkably high activities. The reaction of (i-Pr)2NBCl215 with 2 equiv of CpLi gave a quantitative yield of 11 as a mixture of doublebond isomers (Scheme 1).16 Deprotonation of 11 with LDA followed by reaction with ZrCl4 in ether gave 8b as colorless crystals in 38% yield. 8b was characterized by NMR and mass spectroscopy, elemental analysis, and X-ray crystallography.17 The molecular structure of 8b is displayed in Figure 1. (10) Braunschweig, H.; von Koblinski, C.; Wang, R. Eur. J. Inorg. Chem. 1999, 69. (11) Also see: Braunschweig, H.; Dirk, R.; Mu¨ller, M.; Nguyen, P.; Resendes, R.; Gates, D. P.; Manners, I. Angew. Chem., Int. Ed. Engl. 1997, 36, 2338. (12) Bazan, G. C.; Rodriguez, G.; Ashe, A. J., III; Al-Ahmad, S.; Mu¨ller, C. J. Am. Chem. Soc. 1996, 118, 2291. (13) Bazan, G. C.; Rodriguez, G.; Ashe, A. J., III; Al-Ahmad, S.; Kampf, J. W. Organometallics 1997, 16, 2492. (14) Also see: Rogers, J. S.; Bazan, G. C.; Sperry, C. K. J. Am. Chem. Soc. 1997, 119, 9305. Rogers, J. S.; Lachicotte, R. J.; Bazan, G. C. J. Am. Chem. Soc. 1999, 121, 1288. (15) Gerrard, W.; Hudson, H. R.; Mooney, E. F. J. Chem. Soc. 1960, 5168.

10.1021/om9901776 CCC: $18.00 © 1999 American Chemical Society Publication on Web 05/18/1999

Communications

Organometallics, Vol. 18, No. 12, 1999 2289 Scheme 1

Encouraged by this facile synthesis, we attempted a similar preparation of 4b. The reaction of indenyl(16) Experimental procedures: (a) Bis(cyclopentadienyl) N,N-diisopropylaminoborane (11). A solution of lithium cyclopentadienide (0.31 g, 2.1 mmol) in THF (10 mL) was added dropwise at -78 °C to a solution of diisopropylaminoboron dichloride (0.39 g, 2.1 mmol) in THF (5 mL) at -78 °C. The mixture was slowly warmed and stirred 15 h at room temperature to give a red solution. After solvent removal, the residue was extracted with pentane (3×), and combined extracts were filtered. Solvent was then removed to give product (0.55 g, 100%) as a yellow syrup: 1H NMR spectrum was rather complicated, which suggested the product was a mixture of isomers. 11B NMR (115 MHz, C6D6): δ 39.8 (major), 30.4 (minor). MS (EI, m/e (intensity)): 241 (M+, 47), 226 (40), 176 (16), 93 (100). (b) N,N-Diisopropylaminoboranediyl bis(1-cyclopentadienyl) zirconium dichloride (8b). A solution of LDA (6.93 mmol) in ether (10 mL) was added to a solution of 11 (0.71 g, 2.95 mmol) in Et2O (15 mL) at -78 °C. The mixture was warmed and stirred for 2 h at room temperature to give a slightly turbid solution. The solution was then added to a suspension of ZrCl4 (0.69 g, 2.95 mmol) in Et2O (15 mL) at -78 °C. The resulting mixture was warmed and stirred 15 h at room temperature to give a yellowish suspension. After solvent was removed, residue was extracted with toluene and the extracts were filtered. The toluene solution was concentrated and cooled to -78 °C to give product (0.50 g, 38%) as colorless crystals: Mp ) 222 °C (dec). 1H NMR (400 MHz, CDCl3): δ 1.32 (d, 12H, J ) 7.0 Hz), 2.92 (m, 2H), 5.65 (t, 4H, J ) 2.4 Hz), 6.80 (t, 4H, J ) 2.4 Hz). 13C NMR (100 MHz, CDCl ): δ 24.6, 49.6, 111.4, 125.7. 11B NMR (115 3 MHz, CDCl3): δ 37.6. HRMS (EI) Calcd for C16H2211B35Cl2NZr: 399.0269, found 399.0272. Anal. Calcd for C16H22BCl2NZr: C, 47.88; H, 5.49; N, 3.49. Found: C, 46.95; H, 5.44; N, 3.12. Single crystals suitable for X-ray diffraction analysis were grown from a mixed solvent of CH2Cl2 and hexane. (c) Bis(1-indenyl) N,N-diisopropylaminoborane (12). A solution of lithium indenide (1.50 g, 12.3 mmol) in THF (40 mL) at -78 °C was added dropwise to a solution of diisopropylaminoboron dichloride (1.10 g, 6.0 mmol) in THF (10 mL) at -78 °C. The mixture was slowly warmed and stirred 15 h at room temperature to give a red solution. After solvent removal, residue was extracted with CH2Cl2 (3×), and the extracts were filtered. Removal of solvent gave product (2.12 g, 100%) as a white solid: 1H NMR spectrum was very complicated, which indicated product as a mixture of double isomers. 11B NMR (115 MHz, C D ): δ 41.4. MS (EI, m/e (intensity)): 341 (M+, 6 6 8), 226 (100). HRMS (EI), Calcd for C24H2811BN: 341.2315, found 341.2310. (d) meso-N,N-Diisopropylaminoboranediyl bis(1-indenyl) zirconium dichloride (meso-4b). A solution of LDA (2.62 mmol) in ether (8 mL) was added to a suspension of 12 (0.39 g, 1.41 mmol) in Et2O (10 mL) at -78 °C. The mixture was warmed and stirred 15 h at room temperature to give an orange suspension. After solvent was removed, the residue was extracted with toluene and filtered. Solvent was then removed to give an orange solid, which consists mostly of a mixture of two stereoisomers (rac/meso, ca. 6:4). Repeated recrystallization (3×) from toluene at -78 °C gave the pure meso isomer (0.08 g, 14%) as an orange solid: Mp ) 250-254 °C (dec). 1H NMR (500 MHz, CDCl3): δ 1.54 (d, 6H, J ) 6.6 Hz), 1.57 (d, 6H, J ) 6.8 Hz), 4.27 (m, 2H, NCH), 5.91 (d, 2H, J ) 3.0 Hz), 6.9 (m, 4H), 7.17 (t, 2H, J ) 7.6 Hz), 7.31 (dd, 2H, J ) 8.3, 3.6 Hz), 7.53 (d, 2H, J ) 8.6 Hz). 13C NMR (75 MHz, CDCl3): δ 24.7, 25.0 (NCHCH3), 49.7, 100.2 (BC), 115.7, 117.2, 125.2, 125.3, 125.6, 125.9, 126.5, 131.4. 11B NMR (115 MHz, CDCl3): δ 39.5. HRMS (EI) Calcd for C24H2611B35Cl2Zr: 499.0582, found 499.0606. (e) rac-Bis(dimethylamido)-N,N-diisopropylaminoboranediyl bis(1-indenyl) zirconium (13). Toluene (15 mL) was added to a mixture of 12 (1.01 g, 2.96 mmol) and Zr(NMe2)4 (0.79 g, 2.96 mmol). The resulting solution was heated to 65 °C and stirred for 2 h to give a bright red solution. The reaction vessel was open to the manifold with a moderate N2 flow. 1H NMR spectroscopic analysis of an aliquot showed that the product consists of two isomers with a ratio of 6.7:1. The solution was filtered, concentrated, and cooled to -78 °C to give the pure rac isomer (0.50 g, 33%) as red crystals. The structure was confirmed by X-ray diffraction analysis on single crystals grown from toluene/hexane at

Figure 1. ORTEP view of 8b, showing the atom-numbering scheme. Hydrogen atoms are omitted for clarity. Selected bond distances (Å) and angles (deg): B-C(1), 1.578(6); B-C(6), 1.584(6); ZrCp(c), 2.20; Cp/Cp, 65.5; Cp(c)-ZrCp(c), 119.6, Cl(1)-Zr-Cl(2), 97.9; C(1)-B-C(6), 103.9(3).

lithium with (i-Pr)2NBCl2 afforded a quantitative yield of bis(indenyl)diisopropylaminoborane 12 as a mixture of double-bond isomers. Reaction of 12 with LDA followed by ZrCl4 gave 4b as a mixture of isomers in the ratio of 4:6. The minor isomer, subsequently identified as meso, could be isolated in 14% yield by repeated recrystallization from toluene. The desired rac isomer was obtained by an adaption of the aminolysis procedure used by Jordan et al. for stereoselective synthesis of pure rac-1b.18 Reaction of 12 with Zr(NMe2)4 in toluene at 65 °C for 2 h afforded rac-13 containing only 13% of the meso isomer. Recrystallization from toluene gave pure rac-13 as red crystals in 33% yield. Reaction of 13 with TMSCl followed by recrystallization afforded pure rac-4b in 83% yield. The molecular structure of 4b is illustrated in Figure 2.19 -20 °C. Mp ) 220 °C (dec). 1H NMR (500 MHz, C6D6): δ 1.20 (d, 6H, J ) 6.8 Hz), 1.27 (d, 6H, J ) 6.6 Hz), 2.61 (s, 12H), 3.86 (m, 2H), 6.04 (d, 2H, J ) 2.9 Hz), 6.74 (m, 4H), 7.00 (t, 2H, J ) 7.6 Hz), 7.37 (dd, 2H, J ) 8.5, 0.9 Hz), 7.52 (d, 2H, J ) 8.5 Hz). 13C NMR (90 MHz, C6D6): δ 24.6, 24.9, 47.9 (NCH3), 49.6, 105.6, 112.4, 123.1, 123.2, 124.0, 124.2, 126.3, 129.1. 11B NMR (115 MHz, C6D6): δ 40.8. HRMS (EI) Calcd for C28H3811BN3Zr: 517.2206, found 517.2217. (f) rac-N,NDiisopropylaminoboranediyl bis(1-indenyl) zirconium dichloride (rac4b). TMSCl (2.0 mL, 15.76 mmol) was added to a solution of 13 (0.50 g, 0.96 mmol) in toluene (35 mL) at room temperature. The solution was stirred for 8 h at room temperature to give an orange suspension. Solvent was removed, and residue was washed with pentane (2×) to give the desired product (0.40 g, 83%) as an orange powder. Single crystals suitable for X-ray structural analysis were grown from mixed solvents of CH2Cl2 and hexane at -20 °C. Mp ) 242 °C (dec). 1H NMR (500 MHz, CDCl3): δ 1.50 (d, 6H, J ) 6.8 Hz), 1.55 (d, 6H, J ) 6.6 Hz), 4.24 (hept, 2H, J ) 7.7 Hz), 5.79 (d, 2H, J ) 3.0 Hz), 6.80 (dd, 2H, J ) 3.2, 0.7 Hz), 7.07 (t, 2H, J ) 7.6 Hz), 7.28 (dd, 2H, J ) 7.0, 0.7 Hz), 7.38 (t, 2H, J ) 8.0 Hz), 7.58 (d, 2H, J ) 8.8 Hz). 13C NMR (75 MHz, CDCl3): δ 24.7, 25.0, 49.6, 98.4 (BC), 113.1, 114.1, 122.0, 123.0, 125.8, 126.4, 127.2, 131.9. 11B NMR (115 MHz, CDCl3): δ 39.2. HRMS (EI) Calcd for C24H2611B35Cl2NZr: 499.0582, found 499.0606. Anal. Calcd for C24H26BNCl2Zr: C, 57.48; H, 5.19; N, 2.79. Found: C, 57.46; H, 5.32; N, 2.68. (17) Crystal data for C16H22BCl2NZr (8b): monoclinic, P21/n, a ) 12.1342(2) Å, b ) 10.4836(1) Å, c ) 14.3648(2) Å, β ) 107.9462(5)°, V ) 1738.44(4) Å3, Z ) 4, Dc ) 1.553 g cm-3, T ) 158(2) K, λ(Mo KR) ) 0.710 73 Å. Data were collected on a Siemens SMART CCD. Final R indices [I > 2σ(I)]: R1 ) 0.0555, wR2 ) 0.1071. R indices (all data): R1 ) 0.0711, wR2 ) 0.1123. (18) (a) Diamond, G. M.; Rodewald, S.; Jordan, R. F. Organometallics 1995, 14, 5. (b) Diamond, G. M.; Jordan, R. F.; Petersen, J. L. Organometallics 1996, 15, 4030. (19) Crystal data for C24H26BCl2NZr (4b): triclinic, P1h , a ) 8.5588(1) Å, b ) 10.3633(2) Å, c ) 13.9398(2) Å, R ) 100.919(1)°, β ) 94.946(1)°, γ ) 108.19(1)°, V ) 1139.26(3) Å3, Z ) 2, Dc ) 1.462 g cm-3, T ) 158(2) K, λ(Mo KR) ) 0.710 73 Å. Data were collected on a Siemens SMART CCD. Final R indices [I > 2σ(I)]: R1 ) 0.0217, wR2 ) 0.0575. R indices (all data): R1 ) 0.0238, wR2 ) 0.0591.

2290 Organometallics, Vol. 18, No. 12, 1999

Communications

and 101° for 6b) are nearly identical. This close structural similarity suggests that any chemical differences must be due to different bonding at boron. When activated by excess MAO as a cocatalyst, complexes 8b and 4b are highly active olefin polymerization catalysts. Under typical conditions 8b and 4b effect the copolymerization of mixtures of 1-octene/ ethylene with activities of 0.7 × 106 g/(mol Zr‚atm) and 17 × 106 g/(mol Zr‚atm), respectively.20 Complexes 8b and 4b effect the polymerization of propylene with activities of 0.64 × 105/(mol Zr‚atm) and 2.2 × 106 g/(mol Zr‚atm), respectively.21 The polypropylene resulting from polymerization by 4b had mp ) 113 °C with triad ratios of 82:12:6 (mm/mr/rr) and thus had a modest isotacticity.22 These catalytic activities are in the range shown by Me2Si-bridged zirconocenes.1,23,24 It seems likely that modifications of the indenyl moieties of 4 will give rise to even more active and selective catalysts.24 Further research on borane-bridged metallocenes is under way in our laboratory. Figure 2. ORTEP view of 4b, showing the atom-numbering scheme. Hydrogen atoms are omitted for clarity. Selected bond distances (Å) and angles (deg): B-C(1), 1.586 (2); B-C(10), 1.589 (2); Zr-Cp(c), 2.216; Cp/Cp, 66.8; C(1)-B-C(10), 104.9 (1).

The crystal structure of 4b closely resembles that of 2b in that both compounds have a C2-symmetrically bridged bis indenyl system about the ZrCl2 group. The most important difference is the larger angle between the two coordinated rings, 67° for 4b vs 62° for 2b, due to the smaller size of the bridging atom. There is a consequent increase in the accessibility of the ZrCl2 group, which is likely to affect the relative catalytic activities. The structures of 4b and 8b show very short B-N distances (1.38 Å), which indicates strong B-N π-bonding. This trigonal π-bonding at boron contrasts with the quarternary coordinate bonding shown by 6b. However the geometries of the remaining parts of 6b and 8b are remarkably similar. The dihedral angles between coordinated Cp rings (65.4° for 8b and 65.8° for 6b) and the corresponding CBC angles (104° for 8b

Acknowledgment. This work was supported by the Dow Chemical Company and the National Science Foundation. We are grateful to the Dow Chemical Company for running the polymerization experiments. Supporting Information Available: X-ray crystallographic data of the structures of compounds of 4b and 8b. This material is available free of charge via the Internet at http://pubs.acs.org. OM9901776 (20) For conditions of ethylene polymerization see: Ashe, A. J., III; Fang, X.; Kampf, J. W. Organometallics 1999, 18, 1363. (21) Conditions for the propylene polymerization were similar except the pressure was 23.3 atm and the temperature was 70 °C. (22) The isotacticity of the polypropylene must derive from the diastereoselectivity of the catalyst. This result implies that the chiral aminoboranediyl bridge of 4b remains intact in the active form of the catalyst. See ref 3a. (23) Precise comparison is not possible since the literature polymerizations were run under different conditions. However, see footnote 24c. (24) (a) Burger, P.; Hortmann, K.; Diebold, J.; Brintzinger, H. H. J. Organomet. Chem. 1991, 417, 9. (b) Hortmann, K.; Brintzinger, H. H. New J. Chem. 1992, 16, 51. (c) Stehling, U.; Diebold, J.; Kirsten, R.; Ro¨ll, W.; Brintzinger, H. H.; Ju¨ngling, S.; Mu¨lhaupt, R.; Langhauser, F. Organometallics 1994, 13, 964.